You are on page 1of 32

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/273146695

An Introduction to Pain Pathways and Pain “Targets”

Article in Progress in Molecular Biology and Translational Science · March 2015


DOI: 10.1016/bs.pmbts.2015.01.003 · Source: PubMed

CITATIONS READS

28 3,725

1 author:

Vaskar Das
Rush University Medical Center
61 PUBLICATIONS 299 CITATIONS

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Underlying Pain Mechanism View project

Validation and Evaluation of Direct and/or Indirect Specificity of Novel Targets in Pain View project

All content following this page was uploaded by Vaskar Das on 13 September 2020.

The user has requested enhancement of the downloaded file.


Provided for non-commercial research and educational use only.
Not for reproduction, distribution or commercial use.

This chapter was originally published in the book Progress in Molecular Biology and Translational
Science, Vol. 131 published by Elsevier, and the attached copy is provided by Elsevier for the author's
benefit and for the benefit of the author's institution, for non-commercial research and educational use
including without limitation use in instruction at your institution, sending it to specific colleagues who
know you, and providing a copy to your institution’s administrator.

All other uses, reproduction and distribution, including without limitation commercial reprints, selling or
licensing copies or access, or posting on open internet sites, your personal or institution’s website or
repository, are prohibited. For exceptions, permission may be sought for such use through Elsevier's
permissions site at:
http://www.elsevier.com/locate/permissionusematerial

From Vaskar Das, An Introduction to Pain Pathways and Pain “Targets”. In: Theodore J. Price and
Gregory Dussor, editors, Progress in Molecular Biology and Translational Science, Vol. 131, Burlington:
Academic Press, 2015, pp. 1-30.
ISBN: 978-0-12-801389-2
© Copyright 2015 Elsevier Inc.
Academic Press
Author's personal copy

CHAPTER ONE

An Introduction to Pain Pathways


and Pain “Targets”
Vaskar Das1
Behavioral
1
and Brain Sciences, The University of Texas at Dallas, Richardson, Texas, USA
Corresponding author: e-mail address: vxd150530@utdallas.edu

Contents
1. An Introduction to Pain and Pain Pathways 2
1.1 Neuropathic pain 5
1.2 Inflammatory Pain 7
2. Ion Channels, Receptors, and Other “Targets” for Persistent Inflammatory or
Neuropathic Pain 9
2.1 Ion channels 9
2.2 Sodium channels 10
2.3 Calcium channels 12
2.4 K+ channels 12
2.5 Receptors 13
2.6 Purinergic receptors 13
2.7 Toll-like receptors 13
2.8 PAR receptors 14
2.9 Glutamate receptors 14
2.10 AMPA receptors 15
2.11 NMDA receptors 15
2.12 Metabotropic glutamate receptors 15
2.13 Opioid receptors 16
2.14 TRPV receptors 16
2.15 Prostaglandin (prostanoid) E2 17
2.16 Pronociceptive neurotransmitters 17
3. Summary 18
Acknowledgments 18
References 18

Abstract
The purpose of this chapter is to provide a brief introduction to the anatomy and phys-
iology of pain pathways from peripheral nociceptors to central nervous system areas
involved in the perception and modulation of pain. This chapter also provides a short
introduction to major types of persistent pain: neuropathic and inflammatory persistent
pain, and gives an overview of some important molecular targets that are thought to
mediate these types of pain. These targets, which include ion channels, receptors, and

Progress in Molecular Biology and Translational Science, Volume 131 # 2015 Elsevier Inc. 1
ISSN 1877-1173 All rights reserved.
http://dx.doi.org/10.1016/bs.pmbts.2015.01.003
Author's personal copy
2 Vaskar Das

some neurotransmitters, are further discussed in the context of their relevance as poten-
tial drug targets for the better treatment of pain in patients with persistent pain. Finally,
this chapter introduces several important concepts in pain research that will be primary
topics for chapters that come later in the book.

1. AN INTRODUCTION TO PAIN AND PAIN PATHWAYS


The International Association for the Study of Pain (IASP) has defined
pain as “an unpleasant sensory and emotional experience associated with
actual or potential tissue damage, or described in terms of such damage.”1
When asked to describe their pain, individuals variously described it in terms
of severity (mild, moderate, severe), duration (acute or chronic), and type
(nociceptive, inflammatory, neuropathic).2
Nociceptive pain is the normal acute pain sensation produced by activa-
tion of nociceptors in skin, viscera, and other internal organs in the absence
of sensitization.3–7 It may occur as a result of mechanical, thermal, or chem-
ical noxious stimulation and is variously described as an aching or throbbing
kind of pain.5,6,8,9 Nociceptive pain comprises four main stages: transduc-
tion (i.e., action at receptors in the periphery), transmission (i.e., action
potentials along axons), perception (i.e., cortical processing of nociceptive
input), and modulation (i.e., engagement of descending circuits).4,10–12
Noxious stimuli are first detected by mechanical, thermal, and chemical
nociceptors found on specialized nerve endings present in skin (cutaneous),
viscera, and other internal or external organs.8,9,13,14 Nociceptive impulses
are transmitted from the periphery to the spinal cord via primary afferent
nerve fibers which may be unmyelinated or myelinated.3,15–20 The central
nervous system (CNS) components of this pathway constitute particular
anatomical connections in the spinal cord, brain stem, thalamus, and cortex
(the “pain pathway”), linking the sensory inflow generated in high threshold
primary afferents with those parts of the CNS responsible for conscious
awareness of painful sensations21 (Fig. 1). Unmyelinated nerve fibers are
small diameter C-fibers with diameters in the range 0.4–1.2 μm.22,23
Myelinated primary afferent nerve fibers are the Að-fibers (2–6 μm diame-
ter), whereas the thinly myelinated nerve fibers are the Aβ-fibers (>10 μm
diameter).23,24 Primary afferent C-fibers and Að-fibers are responsible for
transmission of noxious stimuli whereas Aβ-fibers transmit innocuous,
mechanical stimuli such as touch.21–24 Put simply, nociceptors collect infor-
mation from noxious stimuli which are transmitted by C-fibers and
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 3

Figure 1 Simplified schematic diagram of the pain pathway. Pain begins with detection
of damage or potentially damaging stimuli by nociceptive neurons in the periphery that
can transduce this signal into transmission toward the CNS. The first synapse in this
pathway is in the dorsal horn, where these projection neurons can send pain-related
information onto multiple brain areas. Pain perception occurs in the brain and can
be modulated by different centers in the brain. The brain also sends modulatory inputs
back down to the spinal cord to induce pain modulation.

Að-fibers through the dorsal root ganglia to the superficial laminae I/II of
the dorsal horn of the spinal cord.20,23 Að-fibers transmit impulses from
the dorsal horn to deeper laminae (III–IV) of the spinal cord and onto higher
centers in the brain via the spinothalamic tracts.20 Dorsal horn neurons com-
prise (i) projection neurons, (ii) local interneurons, and (iii) propriospinal
neurons.20,25 Although projection neurons are the primary means for trans-
ferring sensory information from the spinal cord to the brain, they are only a
small fraction of the total number of cells in the dorsal horn.23,26 Many
Author's personal copy
4 Vaskar Das

projection neurons have axons that cross the midline and ascend to multiple
areas of the brain including the thalamus, periaqueductal gray matter, lateral
parabrachial area of the pons, and various parts of the medullary reticular
formation.27 These neurons are also involved in activation of endogenous
descending inhibitory pathways that modulate dorsal horn neurons.26
Activity-dependent synaptic plasticity in the spinal cord that generates
postinjury pain hypersensitivity together with the cellular and molecular
mechanisms responsible for this form of neuronal plasticity are termed
“central sensitization.”21 Neuroplastic changes relating to the function,
chemical profile, or structure of the peripheral nervous system are
encompassed by the term “peripheral sensitization” and encompass changes
in receptor, ion-channel, and neurotransmitter expression levels.28,29
Central sensitization in the spinal cord includes sensitization and disinhibi-
tion mechanisms, and supraspinally there are functional changes such as
enlargement of receptive fields.30,31 In the CNS, there are also changes in
the dynamic interplay between neuronal structures and activated glial
cells,30,32,33 a topic covered in depth in Chapter “Nonneuronal Central
Mechanisms of Pain: Glia and Immune Response” by E. Alfonso
Romero-Sandoval and Sarah Sweitzer.
Following tissue injury and inflammation, vasoactive mediators such as
histamine, substance P (SP), serotonin (5-HT), nitric oxide (NO), prosta-
glandins (PGs), and bradykinin are released which activate nociceptors
resulting in nociception.13 This in turn can induce release of pronociceptive
neurotransmitters such as SP, calcitonin gene-related peptide (CGRP),
dynorphin (Dyn), neurokinin A (NKA), glutamate, adenosine triphosphate
(ATP), NO, PGs, and neurotrophins such as brain-derived neurotropic fac-
tor (BDNF), from primary afferents either in the periphery or at the first syn-
apse in the dorsal horn of the spinal cord.13,20,22,34,35 More recently, the
important role of proinflammatory cytokines (e.g., tumor necrosis factor-
alpha (TNF-α), interleukin-1β, interleukin-18, etc.) in peripheral and cen-
tral sensitization mechanisms associated with persistent pain states has begun
to be appreciated.36
Many C-fibers express transient receptor potential vanilloid 1 (TRPV1)
receptors and hence are sensitive to the vanilloid, capsaicin, which is a high-
affinity ligand for TRPV1 receptors.37 TRPV1-expressing C-fibers may be
further subdivided into two major classes:
(i) those that contain the neuropeptides, SP, and CGRP, express the high-
affinity nerve growth factor (NGF) receptor, TrkA, and are develop-
mentally dependent on NGF,34,38,39 and
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 5

(ii) those that express isolectin B4, (IB-4), the P2X3 purinergic receptor,40
fluoride-resistant acid phosphatase, do not contain SP or CGRP,41 and
are dependent on glial cell line-derived neurotrophic factor (GDNF).34

1.1 Neuropathic pain


The IASP has defined neuropathic pain as “Pain initiated or caused by a pri-
mary lesion or dysfunction in the nervous system.”1,42 Neuropathic pain is
variously described by patients as having one or more of the following qual-
ities: burning, tingling, electric shock like, and stabbing or pins and
needles.5,42,43 The appearance of abnormal sensory signs such as allodynia
in response to innocuous (nonnoxious) stimulation and/or hyperalgesia in
response to noxious stimulation is common.43 When neuropathic pain is
evoked, it may be classified as having dysesthetic, hyperalgesic, or allodynic
properties depending upon the dynamic or static characteristics of the
stimulus.44
In recent years, it has begun to be appreciated that the pathobiology of
various neuropathic pain subtypes may differ.45,46 Hence, multiple research
groups have focussed on developing and validating rodent models of each of
these neuropathic pain conditions.47,48 These more relevant rodent models
of neuropathic pain have considerable potential not only in terms of
unraveling the neurobiology of each of these neuropathic pain subtypes
but also in terms of identifying novel targets for discovery of new efficacious
and well-tolerated analgesics to improve relief of these persistent pain
conditions.43,49
Inflammatory and neuroimmune mechanisms contribute to both
peripheral and central sensitization that underpin the pathobiology of neu-
ropathic pain.50 Following peripheral nerve injury, inflammatory cells
including mast cells, neutrophils, macrophages, and T-lymphocytes contrib-
ute to peripheral sensitization and hyperexcitability of injured and adjacent
noninjured primary afferent nerve fibers.50 In the CNS, activation of glial
cells including microglia and astrocytes leads to the production and secretion
of various proinflammatory mediators that promote neuroimmune activa-
tion and can sensitize the central terminals of primary afferent and
second-order neurons to increase the intensity and duration of pain.50–53
However, rodent models of neuropathic pain allow behavioral pain
responses such as mechanical allodynia in response to application of a non-
noxious stimulus (light pressure) to the hindpaws and/or hyperalgesia in
response to application of noxious stimuli (pressure, heat, cold) to the
Author's personal copy
6 Vaskar Das

hindpaws, to be quantified.43,49 The three most commonly used rodent


models of neuropathic pain involve induction of a unilateral chronic con-
striction injury of the sciatic nerve,54 partial sciatic nerve ligation,55 and
L5 spinal nerve ligation.56 Over the past two decades, these models have
been used to identify numerous neuropathic pain “targets” encompassing
various receptors, ion channels, and enzymes, as well as to assess novel pain
therapeutics in development.43,49
More recently, rodent models of varicella zoster virus-induced neuro-
pathic pain, antiretroviral drug-induced neuropathic pain, cancer
chemotherapy-induced neuropathic pain, bone cancer pain, and multiple
sclerosis-induced neuropathic pain have been developed.43,57,58 It is hoped
that through use of these more sophisticated rodent pain models, it will be
possible to gain an enhanced understanding of the pathobiology of these
conditions. Additionally, it may be possible to identify novel “druggable”
targets for drug discovery aimed at producing novel analgesics with
improved efficacy and reduced adverse-event profiles for improved relief
of these chronic pain conditions in the clinical setting.49,59
Persistent ongoing pain secondary to nerve injury is underpinned by
considerable complexity and plasticity at multiple levels of the neu-
raxis.49,58,60–62 Following peripheral nerve injury, ectopic firing of injured
and uninjured afferents induces neuroplastic changes and “central
sensitization” in the spinal cord and the brain, underpinned by both neuro-
nal and nonneuronal mechanisms49 (Fig. 2). Intensive research over the past
two decades has revealed a large number of receptors, enzymes, and ion
channels as potential novel targets for drug discovery programs aimed at

Figure 2 Simplified diagram of mechanisms of neuropathic pain. Peripheral nerve


injury causes many changes in the function and phenotype of injured fibers, but an
important property for neuropathic pain is the generation of ectopic activity (action
potentials without any stimulus, signified by lightning). This ectopic activity drives spon-
taneous pain and plasticity in the dorsal horn and brain that may underlie clinical fea-
tures like allodynia.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 7

producing new drugs for the relief of neuropathic pain.63–65 Although sev-
eral molecules have entered preclinical and clinical development, very few
have been approved by regulatory agencies for clinical use.66 Hence, this
large unmet medical need is driving research in this field.

1.2 Inflammatory Pain


Inflammatory pain is precipitated by an insult to the integrity of tissues at the
cellular level. One of the characteristic features of inflammatory states is that
normally innocuous stimuli can produce pain.67 Inflammation is classically
associated with pain (dolor), heat (calor), redness (rubor), swelling (tumor),
and loss-of-function (function laesa).67,68 Examples of inflammatory pain
include pain secondary to tissue injury and infection as well as rheumatoid
arthritis.69,70 Following tissue injury, nociceptors in the affected tissue become
sensitized due to the release of proinflammatory mediators from damaged cells
and blood vessels at the site of injury as well as from immune cells that invade
the injured site.71 This topic is covered in detail in Chapter “Peripheral
Scaffolding and Signaling Pathways in Inflammatory Pain by Jeske Nathan.”
Inflammatory mediators including protons, 5HT, histamine, adenosine,
bradykinin, prostaglandin E2 (PGE2), NO, IL-1, TNF-α, interleukin-6
(IL-6), leukemia inhibitory factor, and NGF5,72,73 contribute to nociceptor
sensitization so that innocuous stimuli are detected as painful (allodynia) or
there is an exaggerated response to noxious stimuli (hyperalgesia).34,74 The
central terminals of primary afferent nerve fibers (first-order neurons) are
located in the superficial layers (laminae I/II) of the dorsal horn of the spinal
cord.13 Synaptic input from these terminals to second-order neurons in the
spinal cord transfers information created by action potentials in primary
afferents secondary to peripheral noxious stimuli (depending on intensity
and duration), to the thalamus, and then onto the cerebral cortex in the
brain.13 Synaptic function at the central terminals of first-order neurons is
regulated by neurotransmitter release, primarily involving glutamate, and
neuroactive peptides like substance P and CGRP.60–62,75–78
In inflammatory pain, peripheral inflammation induces a phenotypic
switch in primary sensory neurons to induce a change in their neurochem-
ical character and properties.62 This topic is covered in some depth in Chap-
ters “Translation Control of Chronic Pain by Ohannes K. Melemedjian and
Arkady Khoutorsky,” “Regulation of Gene Expression and Pain States by
Epigenetic Mechanisms by S.M. Géranton and K.K. Tochiki,” and
“Commonalities between Pain and Memory Mechanisms and their Mean-
ing for Understanding Chronic Pain by Theodore J Price and Kufreobong E
Inyang.” In brief, this is underpinned by alterations in transcription and
Author's personal copy
8 Vaskar Das

translation of various receptors and ion channels to induce central sensitiza-


tion by virtue of a change in the level of synaptic input produced by the sen-
sitized afferent nerve fibers62 (Fig. 3). Put simply, continuous inputs from
sensitized nociceptive afferents can activate or trigger central sensitization
that is characterized by a reduced threshold of dorsal horn neurons to nox-
ious stimulation16,79–84 (Fig. 3). There is expansion of the receptive fields of
dorsal horn neurons,85,86 and temporal summation of slow postsynaptic
potentials resulting in a cumulative depolarization and a prolonged after dis-
charge or “wind up” of dorsal horn neurons.15 There is also increased excit-
ability of the flexion reflex in response to peripheral stimulation.82,87 The
neural mechanisms underlying central sensitization involve excitatory amino
acids (EAAs, e.g., mainly L-glutamate) acting at AMPA and NMDA recep-
tors with the net result being persistent activation of the NMDA receptor84
to allow Ca2+ entry into neurons and activation of numerous intracellular
signaling pathways.62
SP-expressing C-fibers and BDNF-expressing dorsal root ganglion
(DRG) neurons both have a significant role in inflammation-induced cen-
tral sensitization after exposure of their peripheral terminals to inflammatory
mediators and NGF released from immune cells.62,88,89 Ectopic firing of
sensitized terminals increase Aβ-mediated synaptic input to superficial dorsal
horn neurons62,90 and induction of cyclooxygenase-2 (COX-2) expression
levels to drive production of PGE2.62,91,92 Retrograde transport of NGF
from the peripheral terminals of C-fibers to the dorsal root ganglia induces
upregulation of TRPV1 expression and activation (phosphorylation) of p38
MAPK.39,93 This in turn leads to upregulated synthesis and release of

Figure 3 Simplified diagram of mechanisms of inflammatory pain. Activation of


immune cells during inflammation leads to the release of inflammatory mediators that
act on nociceptors. Many of these inflammatory mediators from immune cells directly
activate or modulate the activity of nociceptors. This can drive spontaneous pain and
plasticity in the dorsal horn and brain that underlies clinical features including allodynia
and hyperalgesia.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 9

proinflammatory cytokines among which IL-1β and TNF-α contribute to


the development of central sensitization by enhancing excitatory and reduc-
ing inhibitory currents, and by activating induction of COX-2.62,91,94 As a
result, GluR2-containing AMPA receptor activation allows entry of Ca2+
into neurons,62,95 and this mechanism generates as much Ca2+ influx as
occurs with NMDA receptor activation during inflammatory pain.62,95

2. ION CHANNELS, RECEPTORS, AND OTHER “TARGETS”


FOR PERSISTENT INFLAMMATORY OR
NEUROPATHIC PAIN
Intensive research over the past two decades has revealed a vast array
of ion channels, receptors, transporters, and enzymes that are potential
“druggable” targets for use in discovery programs aimed at developing
the next generation of analgesic drugs.66 Examples of “pain targets” for
potential modulation by novel analgesic agents include voltage-gated
sodium channels (Nav1.3, Nav1.7, and Nav1.8),12 voltage-gated potassium
channels (Kv1.4) is the sole Kv1 subunit expressed in smaller diameter sen-
sory neurons96–98 suggesting that homomeric Kv1.4 channels predominate
in Aδ and C-fibers arising from these cells.97 By contrast, larger diameter
neurons associated with mechanoreception and proprioception express
high levels of Kv1.1 and Kv1.2 without Kv1.4 or other Kv1 subunits,
suggesting that heteromers of these subunits predominate on large, myelin-
ated afferent axons that extend from these cells.97,99–102 Additional “pain
targets” include voltage-gated calcium channels (VGCC) (α2δ subunits;
Cav2.2, Cav3.1, Cav3.2, Cav3.3),103 acid-sensing ion channels
(ASICs)104 covered to some extent in Chapter “Meningeal Afferent Signal-
ing and the Pathophysiology of Migraine by Carolina Burgos-Vega, Jamie
Moy and Greg Dussor,” NMDA receptors, TRPV1 receptors, covered in
Chapter “Sensory TRP Channels: The Key Transducers of Nociception
and Pain by Aaron D. Mickle, Andrew J. Shepherd and Durga P.
Mohapatra,” NKA, purinergic receptors, toll-like receptors (TLRs),
protease-activated receptors (PAR) receptors, opioid receptors, the norepi-
nephrine transporter, and cyclooxygenases (COX-1/2).105 Many of these
“pain targets” are described briefly in the following sections and in more
detail throughout this volume, as noted and summarized in Fig. 4.

2.1 Ion channels


Ion channels play a key role in nociception and are involved in sensory
transduction (TRPV1), regulation of neuronal excitability (potassium
Author's personal copy
10 Vaskar Das

Figure 4 Mechanisms of inflammation-induced pain in the periphery. Tissue damage


causes an immune response and the release of inflammatory mediators that act on
nociceptors. As described in the text, peripheral nociceptors are finely tuned to detect
mediators released by immune cells via the expression of receptors that bind these
ligands and the presence of ion channels that are altered by the signaling pathways
downstream of activation of these receptors.

channels), action potential propagation (sodium channels, ATP-gated chan-


nels, ASICs), and presynaptic release of various neurotransmitters (calcium
channels).106

2.2 Sodium channels


Voltage-gated sodium channels are considered a major target for the devel-
opment of novel therapies for improving pain management, as the ectopic
firing of primary afferents is associated with abnormal sodium channel
regulation.107–109
Sodium channels comprise an α-subunit containing a voltage-gated
sodium-selective aqueous pore and one or two smaller ancillary β-subunits.110
Sodium channel subtypes differ in their sensitivity to block by tetrodotoxin
(TTX) with six isoforms being sensitive (TTXs: Nav1.1, Nav1.2, Nav1.3,
Nav1.4, Nav1.6, Nav1.7) to block by nanomolar concentrations of TTX
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 11

and three isoforms being resistant to micromolar concentrations of TTX


(Nav1.5, Nav1.8, Nav1.9).111
Acute inflammatory and neuropathic pain can be attenuated or abolished
by local treatment with sodium channel blockers,112 showing that peripheral
nociceptive input is dependent on the presence of functional voltage-gated
sodium channels. Four voltage-gated sodium channel subtypes (Nav1.3,
Nav1.7, Nav1.8, and Nav1.9) are of greatest interest in pain due to their
selective expression in peripheral nerves.107,113–128 For first-order sensory
neurons, Nav1.8 is expressed by the cell body, peripheral terminals, and cen-
tral terminals within the dorsal horn of the spinal cord.110 Anatomical and
electrophysiological evidence indicates that expression of Nav1.9 is largely
restricted to nociceptive Aδ- and C-fibers.110
Highlighting the importance of Nav1.7 in inflammatory pain (Fig. 4),
levels of expression of Nav1.7 are increased in sensory nerve terminals by
inflammation12,128 and following ablation of Nav1.7 in nociceptive neurons,
inflammatory pain responses are greatly reduced.129 Additionally, dominant
gain-of-function mutations in SCN9A, the gene encoding Nav1.7 to
increase DRG neuron excitability are thought to be causal in two inherited
chronic pain disorders in humans, viz, erythromelalgia, characterized by
burning pain and skin redness in the extremities, and paroxysmal extreme
pain disorder, characterized by skin flushing, rectal, periocular, and peri-
mandibular pain evoked principally by mechanical stimuli.109,130 Addition-
ally, congenital indifference to pain due to rare recessive loss-of-function
mutations in SCN9A mean that although individuals so-affected are of nor-
mal intelligence, they often fail to recognize and report pain in response to
injury or infection which can lead to early mortality.131
Conditional knockout of SCN9A in mice abolished mechanical pain,
inflammatory pain, and reflex withdrawal responses to noxious heat recapit-
ulating the pain-free phenotype in humans with SCN9A loss-of-function
mutations.128 Interestingly, as conditional knockout of SCN9A in both sen-
sory and sympathetic neurons in mice with spinal nerve transection, mark-
edly reduced neuropathic pain behavior in these animals, neuropathic pain
therefore appears to involve interaction between sensory and sympathetic
neurons.128
In neuropathic pain, levels of expression of Nav1.3 are increased in dam-
aged peripheral nerves and this is highly correlated with the appearance of a
rapidly repriming sodium current in small DRG neurons consistent with the
notion that Nav1.3 channels make a key contribution to neuronal hyper-
excitability in neuropathic pain.12,105,110,132–135
Author's personal copy
12 Vaskar Das

2.3 Calcium channels


VGCCs play an important role in synaptic transmission and nociceptive
signaling.103,136–138 N-type (Cav2.2) VGCCs are expressed in high density
in the cell bodies of primary afferents in the dorsal root ganglia and on
presynaptic terminals that form synapses with second-order neurons in the dor-
sal horn.139–144 Cav2.2 channels regulate calcium entry, modulate neurotrans-
mitter release, and lead to changes in sensory nerve excitability for the
modulation of pain.67,145–147 Mice lacking Cav2.2 channels have reduced pain
responses secondary to inflammation and peripheral nerve injury.95,103,148,149
T-type VGCCs (Cav3.1, Cav3.2, and Cav3.3) are expressed on the cell
bodies and peripheral terminals of primary afferents, where they contribute
to initiation of the action potential to regulate neuronal excitability.103,150–152
Multiple studies in rodent pain models have implicated the α2δ1 subunit
of presynaptic calcium channels as having an important role in persistent pain
states.153 This is emphasized by clinical studies showing that the anti-
neuropathic drugs, gabapentin, and pregabalin that are ligands at the α2δ1
subunit, have efficacy for the relief of neuropathic pain.154–157 The mech-
anisms of action of these drugs are still controversial, but their widespread
use for neuropathic pain disorders is highlighted in Chapter “Chronic Pain
Syndromes, Mechanisms, and Current Treatments by Justin Sirianni,
Mohab Ibrahim and Amol Patwardhan.”

2.4 K+ channels
Voltage-gated K+ (Kv) channel subunits are expressed in DRG neurons and
have an important physiological role in the regulation of membrane poten-
tials in excitable tissues including nociceptive neurons.101,102,158–160 The Kv
channel subunit Kv1.4 is the sole Kv1 α subunit expressed in smaller diam-
eter neurons, suggesting that homomeric Kv1.4 channels predominate in Aδ
and C-fibers arising from these cells.97,99,100 Additionally, these neurons are
presumably nociceptors, because they also express the TRPV1 capsaicin
receptor, CGRP, and/or Na+ channel SNS/PN3/Nav1.8.97,99,161–164
However, larger diameter neurons associated with mechanoception and
proprioception express high levels of Kv1.1 and Kv1.2 without Kv1.4 or
other Kv1 α subunits, suggesting that heteromers of these subunits predom-
inate on large, myelinated afferent axons that extend from these
cells.97,99,100,164 As the opening of K+ channels leads to hyperpolarization
of the cell membrane and so decreased nerve cell excitability, several Kv
channels are implicated as possible targets for novel pain therapeutics. For
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 13

example, A-type potassium currents contribute significantly to neuronal


excitability and central sensitization in the dorsal horn of the spinal cord
in inflammatory pain.100,165–169
However, abnormal hyperexcitability of primary sensory neurons plays
an important role in neuropathic pain.164,166 Kv channels regulate neuronal
excitability by affecting the resting membrane potential and influencing the
repolarization and frequency of the action potential and may therefore play a
key role in ectopic activity that develops in peripheral nerves driving neu-
ropathic pain.163,164 Additionally, diabetes primarily reduces Kv channel
activity in medium and large DRG neurons.164 Increased BDNF activity
in these neurons likely contributes to the reduction in Kv channel function
through TrkB receptor stimulation in painful diabetic neuropathy.164

2.5 Receptors
In inflammatory pain, multiple receptor classes located on nociceptors are
modulated by vasoactive mediators released from damaged tissues and
immune cells that invade the inflamed tissues.170

2.6 Purinergic receptors


ATP activates P2X purinergic receptors, especially P2X1, P2X3, or P2X7
receptors, to produce pain.171–173 Currently, a range of preclinical studies
are investigating a role for P2X receptors in pain, inflammation, osteoporo-
sis, multiple sclerosis, spinal cord injury, and bladder dysfunction.173 Some
of these have been progressed into clinical trials for rheumatoid arthritis,
pain, and cough.173
P2X3 receptors, located exclusively on small diameter nociceptive-
fibers, are implicated in inflammatory pain and P2X4 receptors on microglia
in the dorsal horn of the spinal cord are implicated in the pathogenesis of
neuropathic pain.174 Antagonists at the P2X7 receptor also reduce pain
behaviors in rodent models of inflammatory and neuropathic pain, again
highlighting that purinergic glial–neuronal interactions are important mod-
ulators of noxious nociceptive neurotransmission.175 Hence, P2X3, P2X4,
and P2X7 receptors are potential targets for novel therapeutics for the treat-
ment of inflammatory and neuropathic pain conditions.176,177

2.7 Toll-like receptors


Proinflammatory central immune signaling contributes significantly to the
initiation and maintenance of heightened pain states because recent
Author's personal copy
14 Vaskar Das

discoveries have implicated the innate immune system, in particular, pattern


recognition TLRs in triggering these proinflammatory central immune sig-
naling events.178 There is considerable interest in the targeting of TLRs on
immune cells for the prevention and treatment of cancer, infection, inflam-
mation, and autoimmune diseases.179 In neuropathic pain, “TLR4 receptors
on peripheral immune cells (e.g., monocytes/macrophages, dendritic cells,
and immune-related cells such as keratinocytes)”170 as well as “activated
microglia in the CNS” appear to have a key role in the establishment of this
condition.180–183 Acute TLR4 antagonism attenuates neuropathic pain
behavior and potentiates opioid antinociception.184 Hence, TLR4 appears
to be a possible target for therapeutic intervention for relief of neuropathic
pain and for augmenting opioid analgesia. As already noted in an earlier sec-
tion of this literature review, dysregulation of chemokines (Section 1) and
their receptors (Section 2.1; Fig 4), particularly fractalkine and its CX3CR1
receptor, appear to play a key role in neuroimmune signaling that contributes
significantly to the pathobiology of neuropathic pain.185 This target is dis-
cussed in more detail in Chapter “Role of Extracellular Damage-Associated
Molecular Pattern Molecules (DAMPS) as Mediators of Persistent Pain by
Jungo Kato and Camilla I Svensson.”

2.8 PAR receptors


PARs are G-protein coupled receptors (GPCRs) that have a unique activa-
tion mechanism involving specific proteolytic cleavage of the amino-terminal
sequence by serine proteases.186–188 PAR1 is expressed by primary afferent
neurons and can modulate nociception.189 PAR2 is expressed by SP- and
CGRP-containing primary afferents.189 Activation of PAR2 induces the
release of the pronociceptive neurotransmitters, SP, and CGRP from both
peripheral and central terminals of primary DRG neurons.189 PAR4 modu-
lates nociceptive responses in normal and inflammatory conditions such that a
PAR4 agonist alleviated inflammatory pain in rats.159,188,190–196

2.9 Glutamate receptors


Glutamate is the major EAA neurotransmitter in the central nervous system
and is found in at least 70% of sensory neurons in the DRGs.197 It is released
from the central terminals of primary afferents and has an important role in
nociceptive neurotransmission.197 Glutamate acts via two main receptor
classes, iGluRs (ionotropic), and mGluRs (metabotropic) with iGluRs fur-
ther subdivided into AMPA, NMDA, and kainate receptors.72,198 AMPA
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 15

and NMDA receptors are directly coupled to cation-permeable ion channels


whereas metabotropic glutamate receptors (mGluRs) are coupled via
G-proteins to soluble second messengers.53,72,198–200 Brief nociceptive stim-
uli primarily activate AMPA receptors whereas stimuli of more prolonged
duration activate NMDA receptors.35,201 Astroglial cells remove excess glu-
tamate from the extracellular space and express the glutamate reuptake trans-
porters, GLAST/EEAT-1 (excitatory amino acid transporter) and glutamate
transporters EEAT-2/GLT-1.202–207

2.10 AMPA receptors


AMPA receptors have an important role in acute spinal processing of noci-
ceptive and nonnociceptive inputs.208 Activation of AMPA receptors by
glutamate results in potent depolarization of dorsal horn neurons to remove
the Mg2+ block, and hence activate NMDA receptors resulting in calcium
influx and initiation of a cascade of downstream signaling events.209 Because
AMPA receptors also have roles in many other CNS functions,208 they are
generally regarded as being unsuitable targets for development of novel pain
therapeutics.208

2.11 NMDA receptors


NMDA receptors (Fig. 1) are located in the superficial and deeper laminae of
the spinal dorsal horn on the central terminals of primary afferents as well as
on membranes that are postsynaptic to the primary afferent.210 All NMDA
receptors display a certain degree of voltage-dependent Mg2+ block and
marked permeability to Ca2+ after removal of the Mg2+ block
(Fig. 1).198,211–214 Spinal NMDA receptors have an important role in
“central sensitization” in the spinal cord in persistent pain states.215 Although
molecules targeting NMDA receptors have potential for the relief of persis-
tent pains such as neuropathic pain, NMDA receptors are involved in nor-
mal physiological functions and so the first generation of these agents were
hampered by CNS side effects in the analgesic dose range.216

2.12 Metabotropic glutamate receptors


Group I mGluRs in laminae I/II of the spinal dorsal horn play an important
role in the transduction of nociceptive input from C-fibers.217 There are
three mGluR classes containing eight cloned mGluRs and in vivo studies
show that these are not involved in acute nociceptive signaling.218–220
Group I mGluRs (mGlu1 and 5) are implicated in central sensitization
Author's personal copy
16 Vaskar Das

and persistent nociception.221 By contrast, activation of group II mGluRs


(mGlu2/3) alleviates neuropathic and inflammatory pain.221 This topic is
discussed in more detail in Chapter “mGluRs Head to Toe in Pain by
Benedict J. Kolber.”

2.13 Opioid receptors


Opioid receptors are members of the superfamily of 7-transmembrane span-
ning region GPCRs that are coupled to intracellular effectors via
G-proteins, mainly of the inhibitory type, i.e., Gi,o.25,222,223 There are high
densities of opioid receptors in various brain regions, the dorsal horn of the
spinal cord, on peripheral nerve terminals, in peripheral tissues including the
gastrointestinal tract and on immune cells (Fig. 2).198,224,225
In the 1990s, three opioid receptor types, viz, μ (MOP), δ (DOP), and κ
(KOP) were cloned, and more recently multiple splice variants of these
receptors, particularly the MOP receptor, have been identified.222,226–228
The endogenous ligands for opioid receptors include the endomorphins
(highly MOP selective), β-endorphin (equal MOP and DOP selectivity),
met- and leu-enkephalin (more selective at DOP than MOP), and Dyn
(KOP selective).229–231 Opioid agonists activate opioid receptors to produce
potent analgesia by activation of the descending inhibitory system to inhibit
ascending excitatory nociceptive transmission (Fig. 2).224,232,233 Studies
using MOP receptor knockout mice show that the antinociceptive and
other effects of morphine and most other clinically available opioid analge-
sics, are produced secondary to activation of the MOP recep-
tor.222,223,234–239 Studies using rodents show that peripheral opioid
actions are increased in inflammation, suggesting that peripherally selective
opioid analgesics may have benefit as future analgesic agents that are devoid
of CNS side effects.25,222

2.14 TRPV receptors


TRPV1 receptor (Fig. 4) is a ligand-gated nonselective cation channel
expressed on primary afferent sensory neurones that can be activated by
exogenous agents (e.g., capsaicin), endogenous substances (e.g., bradykinin,
ethanol, nicotine, anandamide, and insulin) as well as by heat (>43 °C) and
low pH.240–243 Following activation of TRPV1, there is a rapid increase in
intracellular Ca2+ concentrations resulting in nociceptive signal transduction
via C-fibers to produce pain in humans and pain behaviors in animals.243,244
TRPV1 knockout mice exhibit reduced thermal nociception and a loss of
inflammatory thermal hyperalgesia.242,245 However, both NGF and GDNF
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 17

elicit thermal hyperalgesia during peripheral inflammation via an increase in


TRPV1 expression.246 Expression of these two growth factors follows dif-
ferent time courses and they act on distinctive subpopulations of DRG neu-
rons.246 Intradermal injection of capsaicin and NGF produce heat
hyperalgesia via activation of their respective receptors, viz, TRPV1 and
TrkA on sensory nerve terminals. Moreover, PI3K induces heat
hyperalgesia, possibly by regulating TRPV1 activity, in an ERK-dependent
manner, and the PI3K pathway also appears to play a role that is distinct from
ERK (Fig. 4) by regulating the early onset of inflammatory pain.39,93,247,248
TRP channels are discussed in more detail in Chapter “Sensory TRP Chan-
nels: The Key Transducers of Nociception and Pain by Aaron D. Mickle,
Andrew J. Shepherd and Durga P. Mohapatra.”

2.15 Prostaglandin (prostanoid) E2


Inflammatory pain hypersensitivity is regulated by prostaglandin receptors
(EP1, EP2, EP3, EP4 receptors; Fig. 4).249 At the site of inflammation,
PGE2 sensitizes peripheral nociceptors via activation of EP2 receptors that
are present on the peripheral terminals of high threshold sensory nerve fibers
by reducing the nerve firing threshold and increasing responsiveness, which
is the key phenomenon of peripheral sensitization.249,250
Following tissue injury, the synthesis of PGE2 in the spinal cord91 con-
tributes to central sensitization251 and increased excitability of spinal dorsal
horn neurons.249 NSAIDs inhibit prostaglandin synthesis through non-
selective inhibition of constitutively expressed cyclooxygenase COX-1 as
well as the inducible isoform COX-2.252–254

2.16 Pronociceptive neurotransmitters


2.16.1 Nitric oxide
In the CNS, NO is synthesized primarily from the precursor, L-arginine, by
the enzyme nitric oxide synthase (NOS).255 There are three NOS isoforms,
viz, neuronal (nNOS), endothelial (eNOS), and inducible nitric oxide
synthase (iNOS), with nNOS having a role in the modulation of nociceptive
transmission in the spinal cord.255,256 Following glial cell activation in the
CNS, NO is produced by iNOS, to further sensitize nociceptive neurones
and contribute to the maintenance of central sensitization in persistent pain
conditions.256 NO produced in excess by iNOS and nNOS is implicated in
inflammatory and neuropathic pain, and so iNOS and nNOS inhibitors have
been investigated as potential novel agents for alleviation of these chronic
pain conditions.257
Author's personal copy
18 Vaskar Das

2.16.2 Nerve growth factor


NGF levels increase during inflammation258–260 resulting in sensitization of
primary afferents to noxious thermal, mechanical, and chemical stimuli via
upregulated synthesis of TRPV1 receptors as well as SP, CGRP, and bra-
dykinin receptors261–265 (Fig. 4). Anti-NGF therapeutics have advanced
to clinical trials in humans and have shown broad efficacy but also severe,
albeit rare, side effects. NGF signaling is discussed in more detail in Chapters
“Translation Control of Chronic Pain by Ohannes K. Melemedjian and
Arkady Khoutorsky” and “Commonalities between Pain and Memory
Mechanisms and their Meaning for Understanding Chronic Pain by
Theodore J Price and Kufreobong E Inyang.”

3. SUMMARY
Chronic pain is notoriously difficult to treat and so there is a large
unmet clinical need for new treatments to alleviate pain. Rodent models
of neuropathic and inflammatory pain have had a huge impact on our basic
understanding of pain and pain plasticity, a feature that will be discussed
throughout this volume. These basic science investigations have led to
the development of a broad variety of pain targets, some of which were dis-
cussed briefly here and some of which will be discussed in great detail
throughout the rest of this book. The goal of this line of research is to
develop a new generation of pain therapies with broad efficacy and limited
side effects.

ACKNOWLEDGMENTS
Dr. Das is grateful to Professors Maree Smith at University of Queensland and Theodore
Price at University of Texas at Dallas for editing of the chapter.

REFERENCES
1. Merskey B. Classification of Chronic Pain. Descriptions of Chronic Pain Syndromes and Def-
initions of Pain Terms. 2nd ed. Seattle: IASP; 1994:1–222.
2. Rainville P. Brain mechanisms of pain affect and pain modulation. Curr Opin Neurobiol.
2002;12:195–204.
3. Besson JM, Rivot JP. Spinal interneurones involved in presynaptic controls of sup-
raspinal origin. J Physiol. 1973;230:235–254.
4. Carr DB, Goudas LC. Acute pain. Lancet. 1999;353:2051–2058.
5. Woolf CJ, Mannion RJ. Neuropathic pain: aetiology, symptoms, mechanisms, and
management. Lancet. 1999;353:1959–1964.
6. Besson J. The neurobiology of pain. Lancet. 1999;353:1610–1615.
7. Woolf CJ. Pain: moving from symptom control toward mechanism-specific pharma-
cologic management. Ann Intern Med. 2004;140:441–451.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 19

8. Bessou P, Perl E. Response of cutaneous sensory units with unmyelinated fibers to nox-
ious stimuli. J Neurophysiol. 1969;32:1025–1043.
9. Beitel R, Dubner R. Response of unmyelinated (C) polymodal nociceptors to thermal
stimuli applied to monkey’s face. J Neurophysiol. 1976;39:1160–1175.
10. Mense S, Simons DG, Jon Russell MI. Muscle Pain: Understanding Its Nature, Diagnosis,
and Treatment. Pennsylvania, USA: Wolters Kluwer Health; 2001.
11. Berry PH, CHPN C, Paice JA. Patient care: pain management. In: Core Curriculum for
the Generalist Hospice and Palliative Nurse. Dubuque, IA, USA: Kendall Hunt; 2002:63.
12. Wood JN, Boorman JP, Okuse K, Baker MD. Voltage-gated sodium channels and pain
pathways. J Neurobiol. 2004;61:55–71.
13. Millan MJ. The induction of pain: an integrative review. Prog Neurobiol. 1999;57:1–164.
14. Jesse CR, Savegnago L, Nogueira CW. Effect of a metabotropic glutamate receptor 5
antagonist, MPEP, on the nociceptive response induced by intrathecal injection of
excitatory amino acids, substance P, bradykinin or cytokines in mice. Pharmacol Biochem
Behav. 2008;90:608–613.
15. Mendell LM. Physiological properties of unmyelinated fiber projection to the spinal
cord. Exp Neurol. 1966;16:316–332.
16. Price DD, Hull CD, Buchwald NA. Intracellular responses of dorsal horn cells to cuta-
neous and sural nerve A and C fiber stimuli. Exp Neurol. 1971;33:291–309.
17. Gregor M, Zimmermann M. Characteristics of spinal neurones responding to cutane-
ous myelinated and unmyelinated fibres. J Physiol. 1972;221:555–576.
18. Handwerker H, Iggo A, Zimmermann M. Segmental and supraspinal actions on dorsal
horn neurons responding to noxious and non-noxious skin stimuli. Pain.
1975;1:147–165.
19. Calvillo O, Madrid J, Rudomin P. Presynaptic depolarization of unmyelinated primary
afferent fibers in the spinal cord of the cat. Neuroscience. 1982;7:1389–1409.
20. Willis Jr WD, Coggeshall RE. Sensory Mechanisms of the Spinal Cord. Primary Afferent
Neurons and the Spinal Dorsal Horn; Vol. 1. New York, USA: Springer-Verlag; 2004.
21. Woolf CJ. Central sensitization: implications for the diagnosis and treatment of pain.
Pain. 2011;152:S2–S15.
22. Millan MJ. Descending control of pain. Prog Neurobiol. 2002;66:355–474.
23. Urch C. Normal pain transmission. Rev Pain. 2007;1:2–6.
24. Millan MJ. Evidence that an alpha 2A-adrenoceptor subtype mediates antinociception
in mice. Eur J Pharmacol. 1992;215:355–356.
25. Smith MT, South SM. Pain Pharmacology and Analgesia. In: Majewski H, ed. Oxford,
UK: EOLSS Publishers; 2008:1–80.
26. Woolf C, Salter M. Plasticity and pain: role of the dorsal horn. In: Wall and Melzack’s
Textbook of Pain. 5th ed. Philadelphia, USA: Elsevier; 2006:91–105.
27. Todd AJ, Koerber HR. Neuroanatomical substrates of spinal nociception. In: Wall and
Melzack’s Textbook of Pain. Philadelphia, USA: Elsevier; 2006:73–90.
28. Bartlett S, Bonci A, Haass-Koffler C, Naeemuddin M. Ion channel compositions and
uses thereof. 2010. WO Patent App. PCT/US2010/049,149.
29. Hicks Jr EC, Dutton CT. Methods for chronic pain management and treatment using
HCG. 2012. US Patent 20,120,265,712.
30. May A. Chronic pain may change the structure of the brain. Pain. 2008;137:7–15.
31. Carlton SM, Du J, Tan HY, et al. Peripheral and central sensitization in remote spinal
cord regions contribute to central neuropathic pain after spinal cord injury. Pain.
2009;147:265–276.
32. Joksimovic M, Yun BA, Kittappa R, et al. Wnt antagonism of Shh facilitates midbrain
floor plate neurogenesis. Nat Neurosci. 2009;12:125–131.
33. Yong VW, Marks S. The interplay between the immune and central nervous systems in
neuronal injury. Neurology. 2010;74:S9–S16.
Author's personal copy
20 Vaskar Das

34. Dubner R. The neurobiology of persistent pain and its clinical implications. Suppl Clin
Neurophysiol. 2004;57:3–7.
35. Yaksh TL. Calcium channels as therapeutic targets in neuropathic pain. J Pain. 2006;7:
S13–S30.
36. Griffis CA, Compton P, Doering L. The effect of pain on leukocyte cellular adhesion
molecules. Biol Res Nurs. 2006;7:297–312.
37. Rami HK, Gunthorpe MJ. The therapeutic potential of TRPV1 (VR1) antagonists:
clinical answers await. Drug Discov Today. 2004;1:97–104.
38. Basbaum AI. Spinal mechanisms of acute and persistent pain. Reg Anesth Pain Med.
1999;24:59–67.
39. Smith MT, Woodruff TM, Wyse BD, Muralidharan A, Walther T. A small molecule
angiotensin II type 2 receptor (AT2R) antagonist produces analgesia in a Rat model
of neuropathic pain by inhibition of p38 mitogen-activated protein kinase (MAPK)
and p44/p42 MAPK activation in the dorsal root ganglia. Pain Med.
2013;14(10):1557–1568.
40. Vulchanova L, Riedl MS, Shuster SJ, et al. P2X3 is expressed by DRG neurons that
terminate in inner lamina II. Eur J Neurosci. 1998;10:3470–3478.
41. Silverman JD, Kruger L. Acid phosphatase as a selective marker for a class of small sen-
sory ganglion cells in several mammals: spinal cord distribution, histochemical proper-
ties, and relation to fluoride-resistant acid phosphatase (FRAP) of rodents. Somatosens
Mot Res. 1988;5:219–246.
42. Dworkin RH. An overview of neuropathic pain: syndromes, symptoms, signs, and sev-
eral mechanisms. Clin J Pain. 2002;18:343–349.
43. Bridges D, Thompson S, Rice A. Mechanisms of neuropathic pain. Br J Anaesth.
2001;87:12–26.
44. Rasmussen PV, Sindrup SH, Jensen TS, Bach FW. Symptoms and signs in patients with
suspected neuropathic pain. Pain. 2004;110:461–469.
45. Maier C, Baron R, T€ olle T, et al. Quantitative sensory testing in the German Research
Network on Neuropathic Pain (DFNS): somatosensory abnormalities in 1236 patients
with different neuropathic pain syndromes. Pain. 2010;150:439–450.
46. Jensen TS, Baron R, Haanpää M, et al. A new definition of neuropathic pain. Pain.
2011;152:2204–2205.
47. Walker K, Fox AJ, Urban LA. Animal models for pain research. Mol Med Today.
1999;5:319–321.
48. Mogil JS. Animal models of pain: progress and challenges. Nat Rev Neurosci.
2009;10:283–294.
49. Campbell JN, Meyer RA. Mechanisms of neuropathic pain. Neuron.
2006;52:77–92.
50. Moalem G, Tracey DJ. Immune and inflammatory mechanisms in neuropathic pain.
Brain Res Rev. 2006;51:240–264.
51. Basbaum AI, Jessell TM. The perception of pain. In: Kandel ER, Schwartz JH,
Jessell TM, eds. Principles of Neural Science. New York: McGraw-Hill; 2000:472–491.
52. Saadé NE, Massaad CA, Kanaan SA, Safieh-Garabedian B, Jabbur SJ, Atweh SF. Pain
and neurogenic inflammation: a neural substrate for neuroendocrine-immune interac-
tions. In: Saadé NE, Apkarian AV, Jabbur SJ, eds. Pain and Neuroimmune Interactions.
Springer; 2000:111–123.
53. Block BM, Hurley RW, Raja SN. Mechanism-based therapies for pain. Drug News Per-
spect. 2004;17:172–186.
54. Bennett GJ, Xie Y-K. A peripheral mononeuropathy in rat that produces disorders of
pain sensation like those seen in man. Pain. 1988;33:87–107.
55. Seltzer Z, Dubner R, Shir Y. A novel behavioral model of neuropathic pain disorders
produced in rats by partial sciatic nerve injury. Pain. 1990;43:205–218.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 21

56. Kim SH, Chung JM. An experimental model for peripheral neuropathy produced by
segmental spinal nerve ligation in the rat. Pain. 1992;50:355–363.
57. Hasnie F, Breuer J, Parker S, et al. Further characterization of a rat model of varicella
zoster virus–associated pain: relationship between mechanical hypersensitivity and
anxiety-related behavior, and the influence of analgesic drugs. Neuroscience.
2007;144:1495–1508.
58. Jaggi AS, Jain V, Singh N. Animal models of neuropathic pain. Fundam Clin Pharmacol.
2011;25:1–28.
59. Burgess G, Williams D. The discovery and development of analgesics: new mecha-
nisms, new modalities. J Clin Invest. 2010;120:3753.
60. Woolf CJ, Salter MW. Neuronal plasticity: increasing the gain in pain. Science.
2000;288:1765–1769.
61. Ji R-R, Kohno T, Moore KA, Woolf CJ. Central sensitization and LTP: do pain and
memory share similar mechanisms? Trends Neurosci. 2003;26:696–705.
62. Latremoliere A, Woolf CJ. Central sensitization: a generator of pain hypersensitivity by
central neural plasticity. J Pain. 2009;10:895–926.
63. Searle R, Simpson K. Management of chronic pain: current and future options. Fut
Prescriber. 2010;11:11–14.
64. Inoue K, Tsuda M. P2X4 receptors of microglia in neuropathic pain. CNS Neurol Dis-
ord Drug Targets. 2012;11:699–704.
65. Kulkantrakorn K. Emerging concepts and treatment in neuropathic pain. Neurol Asia.
2012;17:265–271.
66. Rice AS, Cimino-Brown D, Eisenach JC, Kontinen VK, Lacroix-Fralish ML, Machin I.
Animal models and the prediction of efficacy in clinical trials of analgesic drugs: a critical
appraisal and call for uniform reporting standards. Pain. 2008;139:243–247.
67. Kidd BL, Urban LA. Mechanisms of inflammatory pain. Br J Anaesth. 2001;87:3–11.
68. Marchand F, Perretti M, McMahon SB. Role of the immune system in chronic pain.
Nat Rev Neurosci. 2005;6:521–532.
69. Tracey KJ. The inflammatory reflex. Nature. 2002;420:853–859.
70. Staud R. Fibromyalgia pain: do we know the source? Curr Opin Rheumatol.
2004;16:157–163.
71. Woolf CJ, Ma Q. Nociceptors—noxious stimulus detectors. Neuron. 2007;55:353–364.
72. Smith M. Neurobiology of pain. In: Molecular Pathomechanisms and New Trends in Drug
Research. London: Taylor & Francis; 2003:302–318.
73. Sessle BJ. Peripheral and central mechanisms of orofacial inflammatory pain. Int Rev
Neurobiol. 2011;97:179–206.
74. Smith MT, Wyse BD, Edwards SR. Small molecule angiotensin II type 2 receptor
(AT2R) antagonists as novel analgesics for neuropathic pain: comparative pharmaco-
kinetics, radioligand binding, and efficacy in rats. Pain Med. 2013;14(5):692–705.
75. Woolf CJ, Thompson SW. The induction and maintenance of central sensitization is
dependent on N-methyl-D-aspartic acid receptor activation; implications for the treat-
ment of post-injury pain hypersensitivity states. Pain. 1991;44:293–299.
76. Cervero F. Visceral pain: mechanisms of peripheral and central sensitization. Ann Med.
1995;27:235–239.
77. Cervero F, Laird J. Mechanisms of touch-evoked pain (allodynia): a new model. Pain.
1996;68:13–23.
78. Cervero F. Understanding Pain. Cambridge, MA, USA: The MIT Press; 2012.
79. Perl ER, Kumazawa T, Lynn B, Kenins P. Sensitization of high threshold receptors
with unmyelinated (C) afferent fibers. Prog Brain Res. 1976;43:263–277.
80. Price DD, Hayes RL, Ruda M, Dubner R. Spatial and temporal transformations of
input to spinothalamic tract neurons and their relation to somatic sensations.
J Neurophysiol. 1978;41(4):933–947.
Author's personal copy
22 Vaskar Das

81. Kenshalo Jr DR, Leonard RB, Chung JM, Willis WD. Responses of primate spino-
thalamic neurons to graded and to repeated noxious heat stimuli. J Neurophysiol.
1979;42:1370–1389.
82. Woolf CJ. Evidence for a central component of post-injury pain hypersensitivity.
Nature. 1983;306:686–688.
83. Coderre TJ, Melzack R. Increased pain sensitivity following heat injury involves a cen-
tral mechanism. Behav Brain Res. 1985;15:259–262.
84. Coderre TJ, Melzack R. The contribution of excitatory amino acids to central sensi-
tization and persistent nociception after formalin-induced tissue injury. J Neurosci.
1992;12:3665–3670.
85. Cook AJ, Woolf CJ, Wall PD, McMahon SB. Dynamic receptive field plasticity in rat
spinal cord dorsal horn following C-primary afferent input. Nature. 1987;325:151–153.
86. Hylden JL, Nahin RL, Traub RJ, Dubner R. Expansion of receptive fields of spinal
lamina I projection neurons in rats with unilateral adjuvant-induced inflammation:
the contribution of dorsal horn mechanisms. Pain. 1989;37:229–243.
87. Woolf CJ, Wall PD. Relative effectiveness of C primary afferent fibers of different ori-
gins in evoking a prolonged facilitation of the flexor reflex in the rat. J Neurosci.
1986;6:1433–1442.
88. Mannion RJ, Doubell TP, Coggeshall RE, Woolf CJ. Collateral sprouting of uninjured
primary afferent A-fibers into the superficial dorsal horn of the adult rat spinal cord after
topical capsaicin treatment to the sciatic nerve. J Neurosci. 1996;16:5189–5195.
89. Neumann S, Doubell TP, Leslie T, Woolf CJ. Inflammatory pain hypersensitivity
mediated by phenotypic switch in myelinated primary sensory neurons. Nature.
1996;384:360–364.
90. Baba H, Doubell TP, Woolf CJ. Peripheral inflammation facilitates Abeta fiber-
mediated synaptic input to the substantia gelatinosa of the adult rat spinal cord.
J Neurosci. 1999;19:859–867.
91. Samad TA, Moore KA, Sapirstein A, Billet S, Allchorne A, Poole S. Interleukin-1beta-
mediated induction of Cox-2 in the CNS contributes to inflammatory pain hypersen-
sitivity. Nature. 2001;410:471–475.
92. Vardeh D, Wang D, Costigan M, Lazarus M, Saper CB, Woolf CJ. COX2 in CNS
neural cells mediates mechanical inflammatory pain hypersensitivity in mice. J Clin
Invest. 2009;119:287–294.
93. Ji R-R, Samad TA, Jin S-X, Schmoll R, Woolf CJ. p38 MAPK activation by NGF in
primary sensory neurons after inflammation increases TRPV1 levels and maintains heat
hyperalgesia. Neuron. 2002;36:57–68.
94. Kawasaki Y, Zhang L, Cheng J-K, Ji R-R. Cytokine mechanisms of central sensitiza-
tion: distinct and overlapping role of interleukin-1beta, interleukin-6, and tumor
necrosis factor-alpha in regulating synaptic and neuronal activity in the superficial spinal
cord. J Neurosci. 2008;28:5189–5194.
95. Luo C, Seeburg PH, Sprengel R, Kuner R. Activity-dependent potentiation of cal-
cium signals in spinal sensory networks in inflammatory pain states. Pain.
2008;140:358–367.
96. Cooper EC, Milroy A, Jan YN, Jan LY, Lowenstein DH. Presynaptic localization of
Kv1. 4-containing A-type potassium channels near excitatory synapses in the hippo-
campus. J Neurosci. 1998;18:965–974.
97. Rasband MN, Park EW, Vanderah TW, Lai J, Porreca F, Trimmer JS. Distinct potas-
sium channels on pain-sensing neurons. Proc Natl Acad Sci USA. 2001;98:13373–13378.
98. Chi XX, Nicol GD. Manipulation of the potassium channel Kv1. 1 and its effect on
neuronal excitability in rat sensory neurons. J Neurophysiol. 2007;98:2683–2692.
99. Takeda M, Kitagawa J, Takahashi M, Matsumoto S. Activation of interleukin-1β
receptor suppresses the voltage-gated potassium currents in the small-diameter trigem-
inal ganglion neurons following peripheral inflammation. Pain. 2008;139:594–602.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 23

100. Takeda M, Tanimoto T, Nasu M, Matsumoto S. Temporomandibular joint inflamma-


tion decreases the voltage-gated K+ channel subtype 1.4-immunoreactivity of trigem-
inal ganglion neurons in rats. Eur J Pain. 2008;12:189–195.
101. Takeda M, Matsumoto S, Sessle BJ, Shinoda M, Iwata K. Peripheral and central
mechanisms of trigeminal neuropathic and inflammatory pain. J Oral Biosci.
2011;53:318–329.
102. Takeda M, Tsuboi Y, Kitagawa J, Nakagawa K, Iwata K, Matsumoto S. Potassium
channels as a potential therapeutic target for trigeminal neuropathic and inflammatory
pain. Mol Pain. 2011;7:1–8.
103. Doan L. Voltage-gated calcium channels and pain. Techniques in Regional Anesthesia and
Pain Management. 2010;14:42–47.
104. Gu Q, Lee L-Y. Acid-sensing ion channels and pain. Pharmaceuticals. 2010;3:
1411–1425.
105. McMahon S, Koltzenburg M, Tracey I, Turk DC. Wall and Melzack’s Textbook of Pain:
Expert Consult-Online. Philadelphia, USA: Elsevier Health Science; 2013.
106. Cregg R, Momin A, Rugiero F, Wood JN, Zhao J. Pain channelopathies. J Physiol.
2010;588:1897–1904.
107. Clare JJ, Tate SN, Nobbs M, Romanos MA. Voltage-gated sodium channels as ther-
apeutic targets. Drug Discov Today. 2000;5:506–520.
108. Devor M. Centralization, central sensitization and neuropathic pain. Focus on
“sciatic chronic constriction injury produces cell-type-specific changes in the
electrophysiological properties of rat substantia gelatinosa neurons” J Neurophysiol.
2006;96:522–523.
109. Fischer TZ, Waxman SG. Familial pain syndromes from mutations of the NaV1. 7
sodium channel. Ann N Y Acad Sci. 2010;1184:196–207.
110. Amir R, Argoff CE, Bennett GJ, et al. The role of sodium channels in chronic inflam-
matory and neuropathic pain. J Pain. 2006;7:1–29.
111. Catterall WA, Perez-Reyes E, Snutch TP, Striessnig J. International Union of Pharma-
cology. XLVIII. Nomenclature and structure-function relationships of voltage-gated
calcium channels. Pharmacol Rev. 2005;57:411–425.
112. Bhattacharya A, Wickenden AD, Chaplan SR. Sodium channel blockers for the treat-
ment of neuropathic pain. Neurotherapeutics. 2009;6:663–678.
113. Lifton RP, Gharavi AG, Geller DS. Molecular mechanisms of human hypertension.
Cell. 2001;104:545–556.
114. Hille B. Ion Channels of Excitable Membranes. Sunderland, MA: Sinauer; 2001.
115. Wang S-Y, Wang GK. Voltage-gated sodium channels as primary targets of diverse
lipid-soluble neurotoxins. Cell Signal. 2003;15:151–159.
116. Felix JP, Williams BS, Priest BT, et al. Functional assay of voltage-gated sodium
channels using membrane potential-sensitive dyes. Assay Drug Dev Technol. 2004;2:
260–268.
117. Nassar MA, Levato A, Stirling LC, Wood JN. Neuropathic pain develops normally in
mice lacking both Na(v)1.7 and Na(v)1.8. Mol Pain. 2005;1:24.
118. Garcia ML, Kaczorowski GJ. Potassium channels as targets for therapeutic intervention.
Sci STKE. 2005;2005(302):pe46.
119. Brochu RM, Dick IE, Tarpley JW, et al. Block of peripheral nerve sodium channels
selectively inhibits features of neuropathic pain in rats. Mol Pharmacol. 2006;
69:823–832.
120. Cummins TR, Sheets PL, Waxman SG. The roles of sodium channels in nociception:
implications for mechanisms of pain. Pain. 2007;131:243–257.
121. Dib-Hajj SD, Cummins TR, Black JA, Waxman SG. From genes to pain:
Na < sub > v</sub > 1.7 and human pain disorders. Trends Neurosci. 2007;30:555–563.
Author's personal copy
24 Vaskar Das

122. Dick IE, Brochu RM, Purohit Y, Kaczorowski GJ, Martin WJ, Priest BT. Sodium
channel blockade may contribute to the analgesic efficacy of antidepressants. J Pain.
2007;8:315–324.
123. Kaczorowski GJ, McManus OB, Priest BT, Garcia ML. Ion channels as drug targets:
the next GPCRs. J Gen Physiol. 2008;131:399–405.
124. Schmalhofer WA, Calhoun J, Burrows R, et al. ProTx-II, a selective inhibitor of
NaV1. 7 sodium channels, blocks action potential propagation in nociceptors. Mol
Pharmacol. 2008;74:1476–1484.
125. England S, de Groot MJ. Subtype-selective targeting of voltage-gated sodium channels.
Br J Pharmacol. 2009;158:1413–1425.
126. Dib-Hajj SD, Binshtok AM, Cummins TR, Jarvis MF, Samad T, Zimmermann K.
Voltage-gated sodium channels in pain states: role in pathophysiology and targets for
treatment. Brain Res Rev. 2009;60:65–83.
127. Dib-Hajj SD, Cummins TR, Black JA, Waxman SG. Sodium channels in normal and
pathological pain. Annu Rev Neurosci. 2010;33:325–347.
128. Minett MS, Nassar MA, Clark AK, et al. Distinct Nav1. 7-dependent pain sensations
require different sets of sensory and sympathetic neurons. Nat Commun. 2012;3:791.
129. Nassar MA, Stirling LC, Forlani G, et al. Nociceptor-specific gene deletion reveals a
major role for Nav1.7 (PN1) in acute and inflammatory pain. Proc Natl Acad Sci
USA. 2004;101:12706–12711.
130. Estacion M, Dib-Hajj S, Benke P, et al. NaV1. 7 gain-of-function mutations as a con-
tinuum: A1632E displays physiological changes associated with erythromelalgia and
paroxysmal extreme pain disorder mutations and produces symptoms of both disorders.
J Neurosci. 2008;28:11079–11088.
131. Goldberg Y, MacFarlane J, MacDonald M, et al. Loss-of-function mutations in the
Nav1. 7 gene underlie congenital indifference to pain in multiple human populations.
Clin Genet. 2007;71:311–319.
132. Rogers M, Tang L, Madge DJ, Stevens EB. The role of sodium channels in neuropathic
pain. Semin Cell Dev Biol. 2006;17(5):571–581 [Elsevier].
133. Cheng J-K, Ji R-R. Intracellular signaling in primary sensory neurons and persistent
pain. Neurochem Res. 2008;33:1970–1978.
134. Costigan M, Scholz J, Woolf CJ. Neuropathic pain: a maladaptive response of the ner-
vous system to damage. Annu Rev Neurosci. 2009;32:1.
135. Baron R, Binder A, Wasner G. Neuropathic pain: diagnosis, pathophysiological mech-
anisms, and treatment. Lancet Neurol. 2010;9:807–819.
136. Snutch TP. Targeting chronic and neuropathic pain: the N-type calcium channel
comes of age. NeuroRx. 2005;2:662–670.
137. Altier C, Dale CS, Kisilevsky AE, et al. Differential role of N-type calcium channel
splice isoforms in pain. J Neurosci. 2007;27:6363–6373.
138. Zamponi GW, Lewis RJ, Todorovic SM, Arneric SP, Snutch TP. Role of voltage-
gated calcium channels in ascending pain pathways. Brain Res Rev. 2009;60:84–89.
139. Nowycky MC, Fox AP, Tsien RW. Three types of neuronal calcium channel with
different calcium agonist sensitivity. Nature. 1985;316:440–443.
140. Cao YQ, Mantyh PW, Carlson EJ, Gillespie AM, Epstein CJ, Basbaum AI. Primary
afferent tachykinins are required to experience moderate to intense pain. Nature.
1998;392:390–394.
141. Tsien RW, Wheeler DB. Voltage-gated calcium channels. In: Carafoli E, Klee CB, eds.
Calcium as a Cellular Regulator. New York, USA: Oxford University Press;
1999:171–199.
142. Miljanich G. Ziconotide: neuronal calcium channel blocker for treating severe chronic
pain. Curr Med Chem. 2004;11:3029–3040.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 25

143. Wermeling DP. Ziconotide, an intrathecally administered n-type calcium channel


antagonist for the treatment of chronic pain. Pharmacotherapy. 2005;25:1084–1094.
144. Cao Y-Q. Voltage-gated calcium channels and pain. Pain. 2006;126:5–9.
145. Maggi CA, Tramontana M, Cecconi R, Santicioli P. Neurochemical evidence for the
involvement of N-type calcium channels in transmitter secretion from peripheral end-
ings of sensory nerves in guinea pigs. Neurosci Lett. 1990;114:203–206.
146. Westenbroek RE, Hell JW, Warner C, Dubel SJ, Snutch TP, Catterall WA. Biochem-
ical properties and subcellular distribution of an N-type calcium hannel α1 subunit.
Neuron. 1992;9:1099–1115.
147. Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signal-
ling. Nat Rev Mol Cell Biol. 2000;1:11–21.
148. Saegusa H, Kurihara T, Zong S, et al. Suppression of inflammatory and neuropathic
pain symptoms in mice lacking the N-type Ca2+ channel. The EMBO J. 2001;
20:2349–2356.
149. Luo Z, Calcutt N, Higuera E, et al. Injury type-specific calcium channel α2δ-1 subunit
up-regulation in rat neuropathic pain models correlates with antiallodynic effects of
gabapentin. J Pharmacol Exp Ther. 2002;303:1199–1205.
150. Jevtovic-Todorovic V, Todorovic SM. The role of peripheral T-type calcium channels
in pain transmission. Cell Calcium. 2006;40:197–203.
151. Jagodic MM, Pathirathna S, Nelson MT, et al. Cell-specific alterations of T-type cal-
cium current in painful diabetic neuropathy enhance excitability of sensory neurons.
J Neurosci. 2007;27:3305–3316.
152. Jagodic MM, Pathirathna S, Joksovic PM, Lee W, Nelson MT, Naik AK. Upregulation
of the T-type calcium current in small rat sensory neurons after chronic constrictive
injury of the sciatic nerve. J Neurophysiol. 2008;99:3151–3156.
153. Maneuf Y, Luo Z, Lee K. α2δ and the mechanism of action of gabapentin in the treat-
ment of pain. Semin Cell Dev Biol. 2006;17(5):565–570 [Elsevier].
154. Finnerup NB, Sindrup SH, Bach FW, Johannesen IL, Jensen TS. Lamotrigine in spinal
cord injury pain: a randomized controlled trial. Pain. 2002;96:375–383.
155. Camerino DC, Desaphy J-F. Grand challenge for ion channels: an underexploited
resource for therapeutics. Front Pharmacol. 2010,1:113.
156. Yang C-M, Chen N-C, Shen H-C, et al. Guideline of neuropathic pain treat-
ment and dilemma from neurological point of view. Acta Neurol Taiwan.
2012;21:136–144.
157. Mindruta I, Cobzaru A-M, Bajenaru OA. Overview of neuropathic pain diagnosis and
assessment—an approach based on mechanisms. In: Cyprian Chukwunonye Udeagha
(ED), InTech, 1–23, Neuropathic Pain. Romania: University of Medicine and
Pharmacy of Bucharest; 2012:1.
158. Kim DS, Choi JO, Rim HD, Cho HJ. Downregulation of voltage-gated potassium
channel α gene expression in dorsal root ganglia following chronic constriction injury
of the rat sciatic nerve. Mol Brain Res. 2002;105:146–152.
159. Kim MS, Jo H, Um JY, et al. Agonists of proteinase-activated receptor 2 induce TNF-α
secretion from astrocytoma cells. Cell Biochem Funct. 2002;20:339–345.
160. Binshtok AM. Mechanisms of nociceptive transduction and transmission: a
machinery for pain sensation and tools for selective analgesia. Int Rev Neurobiol.
2011;97:143–177.
161. Varga AW, Yuan L-L, Anderson AE, et al. Calcium–calmodulin-dependent kinase II
modulates Kv4. 2 channel expression and upregulates neuronal A-type potassium cur-
rents. J Neurosci. 2004;24:3643–3654.
162. Tanimoto T, Takeda M, Nasu M, Kadoi J, Matsumoto S. Immunohistochemical
co-expression of carbonic anhydrase II with Kv1. 4 and TRPV1 in rat small-diameter
trigeminal ganglion neurons. Brain Res. 2005;1044:262–265.
Author's personal copy
26 Vaskar Das

163. Tan Z, Donnelly D, LaMotte R. Effects of a chronic compression of the dorsal root
ganglion on voltage-gated Na + and K+ currents in cutaneous afferent neurons.
J Neurophysiol. 2006;95:1115–1123.
164. Cao XH, Byun HS, Chen SR, Cai YQ, Pan HL. Reduction in voltage-gated K +
channel activity in primary sensory neurons in painful diabetic neuropathy: role of
brain-derived neurotrophic factor. J Neurochem. 2010;114:1460–1475.
165. Kajander KC, Wakisaka S, Bennett GJ. Spontaneous discharge originates in the dorsal
root ganglion at the onset of a painful peripheral neuropathy in the rat. Neurosci Lett.
1992;138:225–228.
166. Rudy B, Chow A, Lau D, et al. Contributions of Kv3 channels to neuronal excitability.
Ann N Y Acad Sci. 1999;868:304–343.
167. Watanabe S, Hoffman DA, Migliore M, Johnston D. Dendritic K + channels contrib-
ute to spike-timing dependent long-term potentiation in hippocampal pyramidal neu-
rons. Proc Natl Acad Sci USA. 2002;99:8366–8371.
168. Hu H-J, Carrasquillo Y, Karim F, et al. The kv4. 2 potassium channel subunit is
required for pain plasticity. Neuron. 2006;50:89–100.
169. Hu H-J, Gereau RW. Metabotropic glutamate receptor 5 regulates excitability and
Kv4. 2-containing K+ channels primarily in excitatory neurons of the spinal dorsal
horn. J Neurophysiol. 2011;105:3010–3021.
170. Ren K, Dubner R. Interactions between the immune and nervous systems in pain.
Nature Med. 2010;16:1267–1276.
171. Burnstock G, Wood JN. Purinergic receptors: their role in nociception and primary
afferent neurotransmission. Curr Opin Neurobiol. 1996;6:526–532.
172. Tsuda M, Koizumi S, Kita A, Shigemoto Y, Ueno S, Inoue K. Mechanical allodynia
caused by intraplantar injection of P2X receptor agonist in rats: involvement of hetero-
meric P2X2/3 receptor signaling in capsaicin-insensitive primary afferent neurons.
J Neurosci. 2000;20:RC90.
173. North RA, Jarvis MF. P2X receptors as drug targets. Mol Pharmacol. 2013;83:759–769.
174. Burnstock G. Purinergic receptors and pain. Curr Pharm Des. 2009;15:1717–1735.
175. Donnelly-Roberts D, McGaraughty S, Shieh C-C, Honore P, Jarvis MF. Painful
purinergic receptors. J Pharmacol Exp Ther. 2008;324:409–415.
176. Dray A. Neuropathic pain: emerging treatments. Br J Anaesth. 2008;101:48–58.
177. Gunosewoyo H, Kassiou M. P2X purinergic receptor ligands: recently patented com-
pounds. Expert Opin Ther Pat. 2010;20:625–646.
178. Nicotra L, Loram LC, Watkins LR, Hutchinson MR. Toll-like receptors in chronic
pain. Exp Neurol. 2012;234:316–329.
179. Abadie J. How can the clinical picture guide appropriate laboratory drug testing in the
treatment of pain clinic patients with opioid analgesics? Pain Med. 2012;13:857–859.
180. Lehnardt S, Massillon L, Follett P, et al. Activation of innate immunity in the CNS
triggers neurodegeneration through a Toll-like receptor 4-dependent pathway. Proc
Natl Acad Sci USA. 2003;100:8514–8519.
181. Jou I, Lee JH, Park SY, Yoon HJ, Joe E-H, Park EJ. Gangliosides trigger inflammatory
responses via TLR4 in brain glia. Am J Pathol. 2006;168:1619–1630.
182. Tang S-C, Arumugam TV, Xu X, et al. Pivotal role for neuronal Toll-like receptors in
ischemic brain injury and functional deficits. Proc Natl Acad Sci USA.
2007;104:13798–13803.
183. Buchanan MM, Hutchinson M, Watkins LR, Yin H. Toll-like receptor 4 in CNS
pathologies. J Neurochem. 2010;114:13–27.
184. Hutchinson MR, Bland ST, Johnson KW, Rice KC, Maier SF, Watkins LR. Opioid-
induced glial activation: mechanisms of activation and implications for opioid analgesia,
dependence, and reward. Sci World J. 2007;7:98–111.
185. Clark AK, Malcangio M. Fractalkine/CX3CR1 signaling during neuropathic pain.
Front Cell Neurosci. 2014;8:121.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 27

186. Nguyen C, Coelho A-M, Grady E, et al. Colitis induced by proteinase-activated


receptor-2 agonists is mediated by a neurogenic mechanism. Can J Physiol Pharmacol.
2003;81:920–927.
187. Grant AD, Cottrell GS, Amadesi S, et al. Protease-activated receptor 2 sensitizes the
transient receptor potential vanilloid 4 ion channel to cause mechanical hyperalgesia
in mice. J Physiol. 2007;578:715–733.
188. Vergnolle N. Protease-activated receptors as drug targets in inflammation and pain.
Pharmacol Ther. 2009;123:292–309.
189. Dale C, Vergnolle N. Protease signaling to G protein-coupled receptors: implications
for inflammation and pain. J Recept Signal Transduct Res. 2008;28:29–37.
190. Garavilla LD, Vergnolle N, Young SH, et al. Agonists of proteinase-activated receptor
1 induce plasma extravasation by a neurogenic mechanism. Br J Pharmacol.
2001;133:975–987.
191. Macfarlane SR, Seatter MJ, Kanke T, Hunter GD, Plevin R. Proteinase-activated
receptors. Pharmacol Rev. 2001;53:245–282.
192. Hollenberg MD, Compton SJ. International union of pharmacology. XXVIII.
Proteinase-activated receptors. Pharmacol Rev. 2002;54:203–217.
193. Asfaha S, Brussee V, Chapman K, Zochodne DW, Vergnolle N. Proteinase-activated
receptor-1 agonists attenuate nociception in response to noxious stimuli. Br J Pharmacol.
2002;135:1101–1106.
194. Coelho A-M, Vergnolle N, Guiard B, Fioramonti J, Bruno L. Proteinases and
proteinase-activated receptor 2: a possible role to promote visceral hyperalgesia in rats.
Gastroenterology. 2002;122:1035–1047.
195. Steinhoff M, Stander S, Seeliger S, Ansel JC, Schmelz M, Luger T. Modern aspects of
cutaneous neurogenic inflammation. Arch Dermatol. 2003;139:1479.
196. Asfaha S, Cenac N, Houle S, et al. Protease-activated receptor-4: a novel mechanism of
inflammatory pain modulation. Br J Pharmacol. 2007;150:176–185.
197. Hokfelt T. Neuropeptides in perspective: the last ten years. Neuron. 1991;7:867–879.
198. Dickenson A. NMDA receptor antagonists: interactions with opioids. Acta Anaesthesiol
Scand. 1997;41:112–115.
199. Yaksh TL, Malmberg A. Central pharmacology of nociceptive transmission.
In: Wall PD, Melzack R, eds. Textbook of Pain. 3rd ed. Churchill Livingstone:
Edinburgh; 1994:165–200.
200. Merighi A, Carmignoto G, Gobbo S, Lossi L, Salio C, Maria VA. Neurotrophins in
spinal cord nociceptive pathways. Prog Brain Res. 2004;146:291–321.
201. South SM, Kohno T, Kaspar BK, et al. A conditional deletion of the NR1 subunit of
the NMDA receptor in adult spinal cord dorsal horn reduces NMDA currents and
injury-induced pain. J Neurosci. 2003;23:5031–5040.
202. Riedel G. Function of metabotropic glutamate receptors in learning and memory.
Trends Neurosci. 1996;19:219–224.
203. Hansson E, Ronnback L. Glial-neuronal signaling and astroglial swelling in physiology
and pathology. Adv Exp Med Biol. 2004;559:313–323.
204. Hansson E, R€ onnbäck L. Altered neuronal–glial signaling in glutamatergic transmission
as a unifying mechanism in chronic pain and mental fatigue. Neurochem Res.
2004;29:989–996.
205. Adolph O, K€ oster S, Räth M, et al. Rapid increase of glial glutamate uptake via block-
ade of the protein kinase A pathway. Glia. 2007;55:1699–1707.
206. Jablonski MM, Freeman NE, Orr WE, et al. Genetic pathways regulating glutamate
levels in retinal M€ uller cells. Neurochem Res. 2011;36:594–603.
207. Lane DJ, Lawen A. The glutamate aspartate transporter (GLAST) mediates
L-glutamate-stimulated ascorbate-release via swelling-activated anion channels in cul-
tured neonatal rodent astrocytes. Cell Biochem Biophys. 2013;65:107–119.
Author's personal copy
28 Vaskar Das

208. Garry EM, Fleetwood-Walker SM. A new view on how AMPA receptors and their
interacting proteins mediate neuropathic pain. Pain. 2004;109:210–213.
209. Raouf RK. Functional Regulation of N-methyl-D-Aspartate Receptors by Serine Threonine
Protein Kinases. Canada: University of Toronto; 1997.
210. Randic M, Jiang M, Cerne R. Long-term potentiation and long-term depression of
primary afferent neurotransmission in the rat spinal cord. J Neurosci. 1993;
13:5228–5241.
211. Dickenson AH, Sullivan AF. Evidence for a role of the NMDA receptor in the fre-
quency dependent potentiation of deep rat dorsal horn nociceptive neurones following
C fibre stimulation. Neuropharmacology. 1987;26:1235–1238.
212. Hollmann M, Boulter J, Maron C, et al. Zinc potentiates agonist-induced currents at
certain splice variants of the NMDA receptor. Neuron. 1993;10:943–954.
213. Hollmann M, Heinemann S. Cloned glutamate receptors. Annu Rev Neurosci. 1994;
17:31–108.
214. Hollmann M, Maron C, Heinemann S. N-glycosylation site tagging suggests a three
transmembrane domain topology for the glutamate receptor GluR1. Neuron.
1994;13:1331–1343.
215. Petrenko AB, Yamakura T, Baba H, Shimoji K. The role of N-methyl-D-aspartate
(NMDA) receptors in pain: a review. Anesth Analg. 2003;97:1108–1116.
216. Childers WE, Baudy RB. N-methyl-D-aspartate antagonists and neuropathic pain: the
search for relief. J Med Chem. 2007;50:2557–2562.
217. Fundytus ME, Fisher K, Dray A, Henry JL, Coderre TJ. In vivo antinociceptive activity
of anti-rat mGluR1 and mGluR5 antibodies in rats. Neuroreport. 1998;9:731–735.
218. Schoepp DD, Jane DE, Monn JA. Pharmacological agents acting at subtypes of meta-
botropic glutamate receptors. Neuropharmacology. 1999;38:1431–1476.
219. Schoepp DD. Unveiling the functions of presynaptic metabotropic glutamate receptors
in the central nervous system. J Pharmacol Exp Ther. 2001;299:12–20.
220. Chiechio S, Nicoletti F. Metabotropic glutamate receptors and the control of chronic
pain. Curr Opin Pharmacol. 2012;12:28–34.
221. Chiechio S, Copani A, Zammataro M, Battaglia G, Nicoletti F. Transcriptional regu-
lation of type-2 metabotropic glutamate receptors: an epigenetic path to novel treat-
ments for chronic pain. Trends Pharmacol Sci. 2010;31:153–160.
222. Smith MT. Differences between and combinations of opioids re-visited. Curr Opin
Anesthiol. 2008;21:596–601.
223. Yekkirala AS, Lunzer MM, McCurdy CR, et al. N-naphthoyl-beta-naltrexamine
(NNTA), a highly selective and potent activator of μ/kappa-opioid heteromers. Proc
Natl Acad Sci USA. 2011;108:5098–5103.
224. Fields HL. Pain modulation: expectation, opioid analgesia and virtual pain. Prog Brain
Res. 1999;122:245–253.
225. Fields H. State-dependent opioid control of pain. Nat Rev Neurosci. 2004;5:565–575.
226. Snyder SH, Pasternak GW. Historical review: opioid receptors. Trends Pharmacol Sci.
2003;24:198–205.
227. Pasternak GW. Molecular biology of opioid analgesia. J Pain Symptom Manag.
2005;29:2–9.
228. Pasternak GW. Opioids and their receptors: are we there yet? Neuropharmacology.
2014;76(Part B):198–203.
229. Raynor K, Kong H, Chen Y, et al. Pharmacological characterization of the cloned
kappa-, delta-, and mu-opioid receptors. Mol Pharmacol. 1994;45:330–334.
230. Smith KM. Investigation of Natural Product Scaffolds for the Development of Opioid Receptor
Ligands. USA: Medicinal Chemistry and the Graduate Faculty of the University of
Kansas; 2012.
Author's personal copy
An Introduction to Pain Pathways and Pain “Targets” 29

231. Li Y, Lefever MR, Muthu D, Bidlack JM, Bilsky EJ, Polt R. Opioid glycopeptide anal-
gesics derived from endogenous enkephalins and endorphins. Future Med Chem.
2012;4:205–226.
232. Yaksh T. Central pharmacology of nociceptive transmission. In: Wall PD, Melzack R,
eds. Textbook of Pain. Edinburg, UK: Churchill Livingstone; 1999:253–308.
233. Whiteside GT, Boulet JM, Walker K. The role of central and peripheral μ opioid
receptors in inflammatory pain and edema: a study using morphine and DiPOA ([8-
(3, 3-diphenyl-propyl)-4-oxo-1-phenyl-1, 3, 8-triaza-spiro [4.5] dec-3-yl]-acetic
acid). J Pharmacol Exp Ther. 2005;314:1234–1240.
234. Hargreaves KM, Pardridge WM. Neutral amino acid transport at the human blood-
brain barrier. J Biol Chem. 1988;263:19392–19397.
235. Joris J, Costello A, Dubner R, Hargreaves KM. Opiates suppress carrageenan-induced
edema and hyperthermia at doses that inhibit hyperalgesia. Pain. 1990;43:95–9103.
236. Zhu Y, King MA, Schuller AG, et al. Retention of supraspinal delta-like analgesia and
loss of morphine tolerance in < i> δ opioid receptor </i > knockout mice. Neuron.
1999;24:243–252.
237. Kieffer BL, Gavériaux-Ruff C. Exploring the opioid system by gene knockout. Prog
Neurobiol. 2002;66:285–306.
238. Hack SP, Bagley EE, Chieng BC, Christie MJ. Induction of δ-opioid receptor function
in the midbrain after chronic morphine treatment. J Neurosci. 2005;25:3192–3198.
239. Gaveriaux-Ruff C, Karchewski LA, Hever X, Matifas A, Kieffer BL. Inflammatory
pain is enhanced in delta opioid receptor-knockout mice. Eur J Neurosci.
2008;27:2558–2567.
240. Tominaga M, Caterina MJ, Malmberg AB, et al. The cloned capsaicin receptor inte-
grates multiple pain-producing stimuli. Neuron. 1998;21:531–543.
241. Szallasi A, Szabo T, Biro T, Modarres S, Blumberg PM, Krause JE. Resiniferatoxin-
type phorboid vanilloids display capsaicin-like selectivity at native vanilloid receptors
on rat DRG neurons and at the cloned vanilloid receptor VR1. Br J Pharmacol.
1999;128:428–434.
242. Caterina MJ, Leffler A, Malmberg AB, et al. Impaired nociception and pain sensation in
mice lacking the capsaicin receptor. Science. 2000;288:306–313.
243. Cortright DN, Szallasi A. Biochemical pharmacology of the vanilloid receptor TRPV1.
An update. Eur J Biochem. 2004;271:1814–1819.
244. Cui M, Honore P, Zhong C, et al. TRPV1 receptors in the CNS play a key role in
broad-spectrum analgesia of TRPV1 antagonists. J Neurosci. 2006;26:9385–9393.
245. Davis JB, Gray J, Gunthorpe MJ, et al. Vanilloid receptor-1 is essential for inflammatory
thermal hyperalgesia. Nature. 2000;405:183–187.
246. Amaya F, Shimosato G, Nagano M, et al. NGF and GDNF differentially regulate
TRPV1 expression that contributes to development of inflammatory thermal
hyperalgesia. Eur J Neurosci. 2004;20:2303–2310.
247. Zhuang Z-Y, Xu H, Clapham DE, Ji R-R. Phosphatidylinositol 3-kinase activates
ERK in primary sensory neurons and mediates inflammatory heat hyperalgesia through
TRPV1 sensitization. J Neurosci. 2004;24:8300–8309.
248. Costa B, Comelli F, Bettoni I, Colleoni M, Giagnoni G. The endogenous fatty acid
amide, palmitoylethanolamide, has anti-allodynic and anti-hyperalgesic effects in a
murine model of neuropathic pain: involvement of CB < sub > 1</sub > TRPV1
and PPARγ receptors and neurotrophic factors. Pain. 2008;139:541–550.
249. Lin C-R, Amaya F, Barrett L, et al. Prostaglandin E2 receptor EP4 contributes to
inflammatory pain hypersensitivity. J Pharmacol Exp Ther. 2006;319:1096–1103.
250. Omote K, Kawamata T, Nakayama Y, Yamamoto H, Kawamata M, Namiki A. Effects
of a novel selective agonist for prostaglandin receptor subtype EP4 on hyperalgesia and
inflammation in monoarthritic model. Anesthesiology. 2002;97:170–176.
Author's personal copy
30 Vaskar Das

251. Minami T, Nakano H, Kobayashi T, et al. Characterization of EP receptor subtypes


responsible for prostaglandin E2-induced pain responses by use of EP1 and EP3 recep-
tor knockout mice. Br J Pharmacol. 2001;133:438–444.
252. Zhang Y, Shaffer A, Portanova J, Seibert K, Isakson PC. Inhibition of
cyclooxygenase-2 rapidly reverses inflammatory hyperalgesia and prostaglandin E2 pro-
duction. J Pharmacol Exp Ther. 1997;283:1069–1075.
253. Turini ME, DuBois RN. Cyclooxygenase-2: a therapeutic target. Annu Rev Med.
2002;53:35–57.
254. Rao P, Knaus EE. Evolution of nonsteroidal anti-inflammatory drugs (NSAIDs):
cyclooxygenase (COX) inhibition and beyond. J Pharm Pharm Sci. 2008;11:81s–110s.
255. Freire MAM, Guimarães JS, Leal WG, Pereira A. Pain modulation by nitric oxide in
the spinal cord. Front Neurosci. 2009;3:175.
256. Covey WC, Ignatowski TA, Renauld AE, Knight PR, Nader ND, Spengler RN.
Expression of neuron-associated tumor necrosis factor alpha in the brain is increased
during persistent pain. Reg Anesth Pain Med. 2002;27:357–366.
257. Miclescu A, Gordh T. Nitric oxide and pain: ‘something old, something new’. Acta
Anaesthesiol Scand. 2009;53:1107–1120.
258. Shu X, Mendell LM. Nerve growth factor acutely sensitizes the response of adult rat
sensory neurons to capsaicin. Neurosci Lett. 1999;274:159–162.
259. Zhu W, Galoyan SM, Petruska JC, Oxford GS, Mendell LM. A developmental switch
in acute sensitization of small dorsal root ganglion (DRG) neurons to capsaicin or nox-
ious heating by NGF. J Neurophysiol. 2004;92:3148–3152.
260. Malin SA, Molliver DC, Koerber HR, et al. Glial cell line-derived neurotrophic factor
family members sensitize nociceptors in vitro and produce thermal hyperalgesia in vivo.
J Neurosci. 2006;26:8588–8599.
261. Okuse K. Pain signalling pathways: from cytokines to ion channels. Int J Biochem Cell
Biol. 2007;39:490–496.
262. Basbaum AI, Bautista DM, Scherrer G, Julius D. Cellular and molecular mechanisms of
pain. Cell. 2009;139:267–284.
263. Premkumar LS. Targeting TRPV1 as an alternative approach to narcotic analgesics to
treat chronic pain conditions. AAPS J. 2010;12:361–370.
264. Schumacher MA. Transient receptor potential channels in pain and inflammation: ther-
apeutic opportunities. Pain Practice. 2010;10:185–200.
265. Pethő G, Reeh PW. Sensory and signaling mechanisms of bradykinin, eicosanoids,
platelet-activating factor, and nitric oxide in peripheral nociceptors. Physiol Rev.
2012;92:1699–1775.

View publication stats

You might also like